Tải bản đầy đủ (.pdf) (10 trang)

Báo cáo khoa học: "Hierarchical Bayesian Language Modelling for the Linguistically Informed" pptx

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (224.62 KB, 10 trang )

Proceedings of the EACL 2012 Student Research Workshop, pages 64–73,
Avignon, France, 26 April 2012.
c
2012 Association for Computational Linguistics
Hierarchical Bayesian Language Modelling
for the Linguistically Informed
Jan A. Botha
Department of Computer Science
University of Oxford, UK

Abstract
In this work I address the challenge of aug-
menting n-gram language models accord-
ing to prior linguistic intuitions. I argue
that the family of hierarchical Pitman-Yor
language models is an attractive vehicle
through which to address the problem, and
demonstrate the approach by proposing a
model for German compounds. In an em-
pirical evaluation, the model outperforms
the Kneser-Ney model in terms of perplex-
ity, and achieves preliminary improvements
in English-German translation.
1 Introduction
The importance of effective language models in
machine translation (MT) and automatic speech
recognition (ASR) is widely recognised. n-gram
models, in particular ones using Kneser-Ney
(KN) smoothing, have become the standard
workhorse for these tasks. These models are not
ideal for languages that have relatively free word


order and/or complex morphology. The ability to
encode additional linguistic intuitions into models
that already have certain attractive properties is an
important piece of the puzzle of improving ma-
chine translation quality for those languages. But
despite their widespread use, KN n-gram mod-
els are not easily extensible with additional model
components that target particular linguistic phe-
nomena.
I argue in this paper that the family of hierarchi-
cal Pitman-Yor language models (HPYLM) (Teh,
2006; Goldwater et al., 2006) are suitable for
investigations into more linguistically-informed
n-gram language models. Firstly, the flexibility
to specify arbitrary back-off distributions makes it
easy to incorporate multiple models into a larger
n-gram model. Secondly, the Pitman-Yor process
prior (Pitman and Yor, 1997) generates distribu-
tions that are well-suited to a variety of power-
law behaviours, as is often observed in language.
Catering for a variety of those is important since
the frequency distributions of, say, suffixes, could
be quite different from that of words. KN smooth-
ing is less flexibility in this regard. And thirdly,
the basic inference algorithms have been paral-
lelised (Huang and Renals, 2009), which should
in principle allow the approach to still scale to
large data sizes.
As a test bed, I consider compounding in Ger-
man, a common phenomenon that creates chal-

lenges for machine translation into German.
2 Background and Related Work
n-gram language models assign probabilities to
word sequences. Their key approximation is that
a word is assumed to be fully determined by n−1
words preceding it, which keeps the number of in-
dependent probabilities to estimate in a range that
is computationally attractive. This basic model
structure, largely devoid of syntactic insight, is
surprisingly effective at biasing MT and ASR sys-
tems toward more fluent output, given a suitable
choice of target language.
But the real challenge in constructing n-gram
models, as in many other probabilistic settings, is
how to do smoothing, since the vast majority of
linguistically plausible n-grams will occur rarely
or be absent altogether from a training corpus,
which often renders empirical model estimates
misleading. The general picture is that probability
mass must be shifted away from some events and
redistributed across others.
The method of Kneser and Ney (1995) and
64
its later modified version (Chen and Goodman,
1998) generally perform best at this smoothing,
and are based on the idea that the number of
distinct contexts a word appears in is an impor-
tant factor in determining the probability of that
word. Part of this smoothing involves discount-
ing the counts of n-grams in the training data;

the modified version uses different levels of dis-
counting depending on the frequency of the count.
These methods were designed with surface word
distributions, and are not necessarily suitable for
smoothing distributions of other kinds of surface
units.
Bilmes and Kirchhoff (2003) proposed a more
general framework for n-gram language mod-
elling. Their Factored Language Model (FLM)
views a word as a vector of features, such that a
particular feature value is generated conditional
on some history of preceding feature values. This
allowed the inclusion of n-gram models over se-
quences of elements like PoS tags and semantic
classes. In tandem, they proposed more compli-
cated back-off paths; for example, trigrams can
back-off to two underlying bigram distributions,
one dropping the left-most context word and the
other the right-most. With the right combina-
tion of features and back-off structure they got
good perplexity reductions, and obtained some
improvements in translation quality by applying
these ideas to the smoothing of the bilingual
phrase table (Yang and Kirchhoff, 2006).
My approach has some similarity to the FLM:
both decompose surface word forms into elements
that are generated from unrelated conditional dis-
tributions. They differ predominantly along two
dimensions: the types of decompositions and con-
ditioning possible, and my use of a particular

Bayesian prior for handling smoothing.
In addition to the HPYLM for n-gram lan-
guage modelling (Teh, 2006), models based on
the Pitman-Yor process prior have also been ap-
plied to good effect in word segmentation (Gold-
water et al., 2006; Mochihashi et al., 2009) and
speech recognition (Huang and Renals, 2007;
Neubig et al., 2010). The Graphical Pitman-Yor
process enables branching back-off paths, which
I briefly revisit in §7, and have proved effective
in language model domain-adaptation (Wood and
Teh, 2009). Here, I extend this general line of
inquiry by considering how one might incorpo-
rate linguistically informed sub-models into the
HPYLM framework.
3 Compound Nouns
I focus on compound nouns in this work for two
reasons: Firstly, compounding is in general a very
productive process, and in some languages (in-
cluding German, Swedish and Dutch) they are
written as single orthographic units. This in-
creases data sparsity and creates significant chal-
lenges for NLP systems that use whitespace to
identify their elementary modelling units. A
proper account of compounds in terms of their
component words therefore holds the potential of
improving the performance of such systems.
Secondly, there is a clear linguistic intuition to
exploit: the morphosyntactic properties of these
compounds are often fully determined by the head

component within the compound. For example,
in “Geburtstagskind” (birthday kid), it is “Kind”
that establishes this compound noun as singular
neuter, which determine how it would need to
agree with verbs, articles and adjectives. In the
next section, I propose a model in the suggested
framework that encodes this intuition.
The basic structure of German compounds
comprises a head component, preceded by one or
more modifier components, with optional linker
elements between consecutive components (Gold-
smith and Reutter, 1998).
Examples
• The basic form is just the concatenation of two
nouns
Auto + Unfall = Autounfall (car crash)
• Linker elements are sometimes added be-
tween components
K
¨
uche + Tisch = K
¨
uchentisch (kitchen table)
• Components can undergo stemming during
composition
Schule + Hof = Schulhof (schoolyard)
• The process is potentially recursive
(Geburt + Tag) + Kind = Geburtstag + Kind
= Geburtstagskind (birthday kid)
The process is not limited to using nouns as

components, for example, the numeral in Zwei-
Euro-M
¨
unze (two Euro coin) or the verb “fahren”
(to drive) in Fahrzeug (vehicle). I will treat all
these cases the same.
65
3.1 Fluency amid sparsity
Consider the following example from the training
corpus used in the subsequent evaluations:
de: Die Neuinfektionen
¨
ubersteigen weiterhin die
Behandlungsbem
¨
uhungen.
en: New infections continue to outpace treatment ef-
forts.
The corpus contains numerous other compounds
ending in “infektionen” (16) or “bem
¨
uhungen”
(117). A standard word-based n-gram model
discriminates among those alternatives using as
many independent parameters.
However, we could gauge the approximate syn-
tactic fluency of the sentence almost as well if we
ignore the compound modifiers. Collapsing all
the variants in this way reduces sparsity and yields
better n-gram probability estimates.

To account for the compound modifiers, a sim-
ple approach is to use a reverse n-gram language
model over compound components, without con-
ditioning on the sentential context. Such a model
essentially answers the question, “Given that the
word ends in ‘infektionen’, what modifier(s), if
any, are likely to precede it?” The vast majority of
nouns will never occur in that position, meaning
that the conditional distributions will be sharply
peaked.
mit der Draht·seil·bahn
Figure 1: Intuition for the proposed generative pro-
cess of a compound word: The context generates the
head component, which generates a modifier compo-
nent, which in turn generates another modifier. (Trans-
lation: “with the cable car”)
3.2 Related Work on Compounds
In machine translation and speech recognition,
one approach has been to split compounds as a
preprocessing step and merge them back together
during postprocessing, while using otherwise un-
modified NLP systems. Frequency-based meth-
ods have been used for determining how aggres-
sively to split (Koehn and Knight, 2003), since
the maximal, linguistically correct segmentation
is not necessarily optimal for translation. This
gave rise to slight improvements in machine trans-
lation evaluations (Koehn et al., 2008), with fine-
tuning explored in (Stymne, 2009). Similar ideas
have also been employed for speech recognition

(Berton et al., 1996) and predictive-text input
(Baroni and Matiasek, 2002), where single-token
compounds also pose challenges.
4 Model Description
4.1 HPYLM
Formally speaking, an n-gram model is an
(n − 1)-th order Markov model that approxi-
mates the joint probability of a sequence of
words w as
P (w) ≈
|w|

i=1
P (w
i
|w
i−n+1
, . . . , w
i−1
),
for which I will occasionally abbreviate a con-
text [w
i
, . . . , w
j
] as u. In the HPYLM, the condi-
tional distributions P (w|u) are smoothed by plac-
ing Pitman-Yor process priors (PYP) over them.
The PYP is defined through its base distribution,
and a strength (θ) and discount (d) hyperparame-

ter that control its deviation away from its mean
(which equals the base distribution).
Let G
[u,v]
be the PYP-distributed trigram distri-
bution P (w|u, v). The hierarchy arises by using
as base distribution for the prior of G
[u,v]
another
PYP-distributed G
[v]
, i.e. the distribution P (w|v).
The recursion bottoms out at the unigram distri-
bution G

, which is drawn from a PYP with base
distribution equal to the uniform distribution over
the vocabulary W. The hyperparameters are tied
across all priors with the same context length |u|,
and estimated during training.
G
0
= Uniform(|W|)
G

∼ P Y (d
0
, θ
0
, G

0
)
.
.
.
G
π(u)
∼ P Y (d
|π(u)|
, θ
|π(u)|
, G
π(π(u))
)
G
u
∼ P Y (d
|u|
, θ
|u|
, G
π(u)
)
w ∼ G
u
,
where π(u) truncates the context u by dropping
the left-most word in it.
4.2 HPYLM+c
Define a compound word ˜w as a sequence of

components [c
1
, . . . , c
k
], plus a sentinel symbol $
marking either the left or the right boundary of the
word, depending on the direction of the model. To
maintain generality over this choice of direction,
66
let Λ be an index set over the positions, such that
c
Λ
1
always designates the head component.
Following the motivation in §3.1, I set up the
model to generate the head component c
Λ
1
condi-
tioned on the word context u, while the remaining
components ˜w \ c
Λ
1
are generated by some model
F , independently of u.
To encode this, I modify the HPYLM in two
ways: 1) Replace the support with the reduced vo-
cabulary M, the set of unique components c ob-
tained when segmenting the items in W. 2) Add
an additional level of conditional distributions H

u
(with |u| = n − 1) where items from M combine
to form the observed surface words:
G
u
. . . (as before, except G
0
=Uniform(|M|))
H
u
∼ P Y (d
|u|
, θ
|u|
, G
u
× F )
˜w ∼ H
u
So the base distribution for the prior of the word
n-gram distribution H
u
is the product of a distri-
bution G
u
over compound heads, given the same
context u, and another (n

-gram) language model
F over compound modifiers, conditioned on the

head component.
Choosing F to be a bigram model (n

=2) yields
the following procedure for generating a word:
c
Λ
1
∼ G
u
for i = 2 to k
c
Λ
i
∼ F (·|c
Λ
i−1
)
The linguistically motivated choice for condi-
tioning in F is Λ
ling
= [k, k − 1, . . . , 1] such that
c
Λ
1
is the true head component; $ is drawn from
F (·|c
1
) and marks the left word boundary.
In order to see if the correct linguistic intuition

has any bearing on the model’s extrinsic perfor-
mance, we will also consider the reverse, sup-
posing that the left-most component were actu-
ally more important in this task, and letting the
remaining components be generated left-to-right.
This is expressed by Λ
inv
= [1, . . . , k], where $
this time marks the right word boundary and is
drawn from F (·|c
k
).
To test whether Kneser-Ney smoothing is in-
deed sometimes less appropriate, as conjectured
earlier, I will also compare the case where
F = F
KN
, a KN-smoothed model, with the case
where F = F
HP Y LM
, another HPYLM.
Linker Elements In the preceding definition of
compound segmentation, the linker elements do
not form part of the vocabulary M. Regarding
linker elements as components in their own right
would sacrifice important contextual information
and disrupt the conditionals F (·|c
Λ
i−1
). That is,

given K
¨
uche·n·tisch, we want P (K
¨
uche|Tisch) in
the model, but not P (K
¨
uche|n).
But linker elements need to be accounted
for somehow to have a well-defined generative
model. I follow the pragmatic option of merg-
ing any linkers onto the adjacent component – for
Λ
ling
merging happens onto the preceding compo-
nent, while for Λ
inv
it is onto the succeeding one.
This keeps the ‘head’ component c
Λ
1
in tact.
More involved strategies could be considered,
and it is worth noting that for German the pres-
ence and identity of linker elements between c
i
and c
i+1
are in fact governed by the preceding
component c

i
(Goldsmith and Reutter, 1998).
5 Training
For ease of exposition I describe inference with
reference to the trigram HPYLM+c model with
a bigram HPYLM for F, but the general case
should be clear.
The model is specified by the latent vari-
ables (G
[∅]
, G
[v]
, G
[u,v]
, H
[u,v]
, F

, F
c
), where
u, v ∈ W, c ∈ M, and hyperparameters Ω =
{d
i
, θ
i
} ∪ {d

j
, θ


j
} ∪ {d

2
, θ

2
}, where i = 0, 1, 2,
j = 0, 1, single primes designate the hyperpa-
rameters in F
HP Y LM
and double primes those of
H
[u,v]
. We can construct a collapsed Gibbs sam-
pler by marginalising out these latent variables,
giving rise to a variant of the hierarchical Chinese
Restaurant Process in which it is straightforward
to do inference.
Chinese Restaurant Process A direct repre-
sentation of a random variable G drawn from a
PYP can be obtained from the so-called stick-
breaking construction. But the more indirect rep-
resentation by means of the Chinese Restaurant
Process (CRP) (Pitman, 2002) is more suitable
here since it relates to distributions over items
drawn from such a G. This fits the current set-
ting, where words w are being drawn from a PYP-
distributed G.

Imagine that a corpus is created in two phases:
Firstly, a sequence of blank tokens x
i
is instanti-
ated, and in a second phase lexical identities w
i
are assigned to these tokens, giving rise to the
67
observed corpus. In the CRP metaphor , the se-
quence of tokens x
i
are equated with a sequence
of customers that enter a restaurant one-by-one to
be seated at one of an infinite number of tables.
When a customer sits at an unoccupied table k,
they order a dish φ
k
for the table, but customers
joining an occupied table have to dine on the dish
already served there. The dish φ
i
that each cus-
tomer eats is equated to the lexical identity w
i
of
the corresponding token, and the way in which ta-
bles and dishes are chosen give rise to the charac-
teristic properties of the CRP:
More formally, let x
1

, x
2
, . . . be draws from G,
while t is the number of occupied tables, c the
number of customers in the restaurant, and c
k
the
number of customers at the k-th table. Condi-
tioned on preceding customers x
1
, . . . , x
i−1
and
their arrangement, the i-th customer sits at table
k = k

according to the following probabilities:
Pr(k

| . . . ) ∝

c
k

− d occupied table k

θ + dt unoccupied table t + 1
Ordering a dish for a new table corresponds to
drawing a value φ
k

from the base distribution G
0
,
and it is perfectly acceptable to serve the same
kind of dish at multiple tables.
Some characteristic behaviour of the CRP can
be observed easily from this description: 1) As
more customers join a table, that table becomes
a more likely choice for future customers too.
2) Regardless of how many customers there are,
there is always a non-zero probability of joining
an unoccupied table, and this probability also de-
pends on the number of total tables.
The dish draws can be seen as backing off to
the underlying base distribution G
0
, an important
consideration in the context of the hierarchical
variant of the process explained shortly. Note that
the strength and discount parameters control the
extent to which new dishes are drawn, and thus
the extent of reliance on the base distribution.
The predictive probability of a word w given a
seating arrangement is given by
Pr(w| . . . ) ∝ c
w
− dt
w
+ (θ + dt)G
0

(w)
In smoothing terminology, the first term can be
interpreted as applying a discount of dt
w
to the
observed count c
w
of w; the amount of dis-
count therefore depends on the prevalence of the
word (via t
w
). This is one significant way in
which the PYP/CRP gives more nuanced smooth-
ing than modified Kneser-Ney, which only uses
four different discount levels (Chen and Good-
man, 1998). Similarly, if the seating dynamics
are constrained such that each dish is only served
once (t
w
= 1 for any w), a single discount level
is affected, establishing direct correspondence to
original interpolated Kneser-Ney smoothing (Teh,
2006).
Hierarchical CRP When the prior of G
u
has
a base distribution G
π(u)
that is itself PYP-
distributed, as in the HPYLM, the restaurant

metaphor changes slightly. In general, each node
in the hierarchy has an associated restaurant.
Whenever a new table is opened in some restau-
rant R, another customer is plucked out of thin air
and sent to join the parent restaurant pa(R). This
induces a consistency constraint over the hierar-
chy: the number of tables t
w
in restaurant R must
equal the number of customers c
w
in its parent
pa(R).
In the proposed HPYLM+c model using
F
HP Y LM
, there is a further constraint of a simi-
lar nature: When a new table is opened and serves
dish φ = ˜w in the trigram restaurant for H
[u,v]
,
a customer c
Λ
1
is sent to the corresponding bi-
gram restaurant for G
[u,v]
, and customers c
Λ
2:k

,$
are sent to the restaurants for F
c

, for contexts
c

= c
Λ
1:k−1
. This latter requirement is novel here
compared to the hierarchical CRP used to realise
the original HPYLM.
Sampling Although the CRP allows us to re-
place the priors with seating arrangements S,
those seating arrangements are simply latent vari-
ables that need to be integrated out to get a true
predictive probability of a word:
p(w|D) =

S,Ω
p(w|S, Ω)p(S, Ω|D),
where D is the training data and, as before, Ω are
the parameters. This integral can be approximated
by averaging over m posterior samples (S, Ω)
generated using Markov chain Monte Carlo meth-
ods. The simple form of the conditionals in the
CRP allows us to do a Gibbs update whereby the
table index k of a customer is resampled condi-
tioned on all the other variables. Sampling a new

seating arrangement S for the trigram HPYLM+c
thus corresponds to visiting each customer in the
restaurants for H
[u,v]
, removing them while cas-
cading as necessary to observe the consistency
68
across the hierarchy, and seating them anew at
some table k

.
In the absence of any strong intuitions about ap-
propriate values for the hyperparameters, I place
vague priors over them and use slice sampling
1
(Neal, 2003) to update their values during gener-
ation of the posterior samples:
d ∼ Beta(1, 1) θ ∼ Gamma(1, 1)
Lastly, I make the further approximation of
m = 1, i.e. predictive probabilities are informed
by a single posterior sample (S, Ω).
6 Experiments
The aim of the experiments reported here is to test
whether the richer account of compounds in the
proposed language models has positive effects on
the predictability of unseen text and the genera-
tion of better translations.
6.1 Methods
Data and Tools Standard data preprocessing
steps included normalising punctuation, tokenis-

ing and lowercasing all words. All data sets are
from the WMT11 shared-task.
2
. The full English-
German bitext was filtered to exclude sentences
longer than 50, resulting in 1.7 million parallel
sentences; word alignments were inferred from
this using the Berkeley Aligner (Liang et al.,
2006) and used as basis from which to extract a
Hiero-style synchronous CFG (Chiang, 2007).
The weights of the log-linear translation mod-
els were tuned towards the BLEU metric on
development data using cdec’s (Dyer et al.,
2010) implementation of MERT (Och, 2003).
For this, the set news-test2008 (2051 sen-
tences) was used, while final case-insensitive
BLEU scores are measured on the official test set
newstest2011 (3003 sentences).
All language models were trained on the target
side of the preprocessed bitext containing 38 mil-
lion tokens, and tested on all the German devel-
opment data (i.e. news-test2008,9,10).
Compound segmentation To construct a seg-
mentation dictionary, I used the 1-best segmenta-
tions from a supervised MaxEnt compound split-
ter (Dyer, 2009) run on all token types in bitext. In
addition, word-internal hyphens were also taken
1
Mark Johnson’s implementation, .
brown.edu/

˜
mj/Software.htm
2
/>as segmentation points. Finally, linker elements
were merged onto components as discussed in
§4.2. Any token that is split into more than one
part by this procedure is regarded as a compound.
The effect of the individual steps is summarised
in Table 1.
# Types Example
None 350998 Geburtstagskind
pre-merge 201328 Geburtstag·kind
merge, Λ
ling
150980 Geburtstags·kind
merge, Λ
inv
162722 Geburtstag·skind
Table 1: Effect of segmentation on vocabulary size.
Metrics For intrinsic evaluation of language
models, perplexity is a common metric. Given a
trained model q, the perplexity over the words τ
in unseen test set T is exp


1
|T |

τ
ln(q(τ ))


.
One convenience of this per-word perplexity is
that it can be compared consistently across dif-
ferent test sets regardless of their lengths; its neat
interpretation is another: a model that achieves a
perplexity of η on a test set is on average η-ways
confused about each word. Less confusion and
therefore lower test set perplexity is indicative of
a better model. This allows different models to be
compared relative to the same test set.
The exponent above can be regarded as an
approximation of the cross-entropy between the
model q and a hypothetical model p from which
both the training and test set were putatively gen-
erated. It is sometimes convenient to use this as
an alternative measure.
But a language model only really becomes use-
ful when it allows some extrinsic task to be exe-
cuted better. When that extrinsic task is machine
translation, the translation quality can be assessed
to see if one language model aids it more than an-
other. The obligatory metric for evaluating ma-
chine translation quality is BLEU (Papineni et al.,
2001), a precision based metric that measures how
close the machine output is to a known correct
translation (the reference sentences in the test set).
Higher precision means the translation system is
getting more phrases right.
Better language model perplexities sometimes

lead to improvements in translation quality, but
it is not guaranteed. Moreover, even when real
translation improvements are obtained, they are
69
PPL c-Cross-ent.
mKN 441.32 0.1981
HPYLM 429.17 0.1994
F
KN
Λ
ling
432.95 0.2028
F
KN
Λ
inv
446.84 0.2125
F
HP Y LM
Λ
ling
421.63 0.1987
F
HP Y LM
Λ
inv
435.79 0.2079
Table 2: Monolingual evaluation results. The second
column shows perplexity measured all WMT11 Ger-
man development data (7065 sentences). At the word

level, all are trigram models, while F are bigram mod-
els using the specified segmentation scheme. The third
column has test cross-entropies measured only on the
6099 compounds in the test set (given their contexts ).
not guaranteed to be noticeable in the BLEU
score, especially when targeting an arguably nar-
row phenomenon like compounding.
BLEU
mKN 13.11
HPYLM 13.20
F
HP Y LM
, Λ
ling
13.24
F
HP Y LM
, Λ
inv
13.32
Table 3: Translation results, BLEU (1-ref), 3003 test
sentences. Trigram language models, no count prun-
ing, no “unknown word” token.
P / R / F
mKN 22.0 / 17.3 / 19.4
HPYLM 21.0 / 17.8 / 19.3
F
HP Y LM
, Λ
ling

23.6 / 17.3 / 19.9
F
HP Y LM
, Λ
inv
24.1 / 16.5 / 19.6
Table 4: Precision, Recall and F-score of compound
translations, relative to reference set (72661 tokens, of
which 2649 are compounds).
6.2 Main Results
For the monolingual evaluation, I used an interpo-
lated, modified Kneser-Ney model (mKN) and an
HPYLM as baselines. It has been shown for other
languages that HPYLM tends to outperform mKN
(Okita and Way, 2010), but I am not aware of this
result being demonstrated on German before, as I
do in Table 2.
The main model of interest is HPYLM+c us-
ing the Λ
ling
segmentation and a model F
HP Y LM
over modifiers; this model achieves the lowest
perplexity, 4.4% lower than the mKN baseline.
Next, note that using F
KN
to handle the modi-
fiers does worse than F
HP Y LM
, confirming our

expectation that KN is less appropriate for that
task, although it still does better than the original
mKN baseline.
The models that use the linguistically im-
plausible segmentation scheme Λ
inv
both fare
worse than their counterparts that use the sensible
scheme, but of all tested models only F
KN
& Λ
inv
fails to beat the mKN baseline. This suggests that
in some sense having any account whatsoever of
compound formation tends to have a beneficial ef-
fect on this test set – the richer statistics due to a
smaller vocabulary could be sufficient to explain
this – but to get the most out of it one needs the
superior smoothing over modifiers (provided by
F
HP Y LM
) and adherence to linguistic intuition
(via Λ
ling
).
As for the translation experiments, the rela-
tive qualitative performance of the two baseline
language models carries over to the BLEU score
(HPYLM does 0.09 points better than KN), and is
further improved upon slightly by using two vari-

ants of HPYLM+c (Table 3).
6.3 Analysis
To get a better idea of how the extended mod-
els employ the increased expressiveness, I calcu-
lated the cross-entropy over only the compound
words in the monolingual test set (second column
of Table 2). Among the HPYLM+c variants, we
see that their performance on compounds only is
consistent with their performance (relative to each
other) on the whole corpus. This implies that
the differences in whole-corpus perplexities are at
least in part due to their different levels of adept-
ness at handling compounds, as opposed to some
fluke event.
It is, however, somewhat surprising to observe
that HPYLM+c do not achieve a lower com-
pound cross-entropy than the mKN baseline, as it
suggests that HPYLM+c’s perplexity reductions
compared to mKN arise in part from something
other than compound handling, which is their
whole point.
This discrepancy could be related to the fair-
ness of this direct comparison of models that ul-
70
timately model different sets of things: Accord-
ing to the generative process of HPYLM+c (§4),
there is no limit on the number of components in
a compound: in theory, an arbitrary number of
components c ∈ M can combine to form a word.
HPYLM+c is thus defined over a countably infi-

nite set of words, thereby reserving some prob-
ability mass for items that will never be realised
in any corpus, whereas the baseline models are
defined only over the finite set W. These direct
comparisons are thus lightly skewed in favour of
the baselines. This bolsters confidence in the per-
plexity reductions presented in the previous sec-
tion, but the skew may afflict compounds more
starkly, leading to the slight discrepancy observed
in the compound cross-entropies. What matters
more is the performance among the HPYLM+c
variants, since they are directly comparable.
To home in still further on the compound mod-
elling, I selected those compounds for which
HPYLM+c (F
HP Y LM
, Λ
ling
) does best/worst in
terms of the probabilities assigned, compared to
the mKN baseline (see Table 5). One pattern that
emerges is that the “top” compounds mostly con-
sist of components that are likely to be quite com-
mon, and that this improves estimates both for n-
grams that are very rare (the singleton “senkun-
gen der treibhausgasemmissionen” = decreases in
green house gas emissions) or relatively common
(158, “der hauptstadt” = of the capital).
n-gram ∆ C
gesichts·punkten 0.064 335

700 milliarden us-·dollar 0.021 2
s. der treibhausgas·emissionen 0.018 1
r. der treibhausgas·emissionen 0.011 3
ministerium f
¨
ur land·wirtschaft 0.009 11
bildungs·niveaus 0.009 14
newt ging·rich* -0.257 2
nouri al-·maliki* -0.257 3
klerikers moqtada al-·sadr* -0.258 1
nuri al-·maliki* -0.337 3
sankt peters·burg* -0.413 35
n
¨
achtlichem flug·l
¨
arm -0.454 2
Table 5: Compound n-grams in the test set for which
the absolute difference ∆ = P
HPYLM+c
−P
mKN
is great-
est. C is n-gram count in the training data. Asterisks
denote words that are not compounds, linguistically
speaking. Abbrevs: r. = reduktionen, s.= senkungen
On the other hand, the “bottom” compounds
are mostly ones whose components will be un-
common; in fact, many of them are not truly com-
pounds but artefacts of the somewhat greedy seg-

mentation procedure I used. Alternative proce-
dures will be tested in future work.
Since the BLEU scores do not reveal much
about the new language models’ effect on com-
pound translation, I also calculated compound-
specific accuracies, using precision, recall and
F-score (Table 4). Here, the precision for a
single sentence would be 100% if all the com-
pounds in the output sentence occur in the ref-
erence translation. Compared to the baselines,
the compound precision goes up noticeably under
the HPYLM+c models used in translation, with-
out sacrificing on recall. This suggests that these
models are helping to weed out incorrectly hy-
pothesised compounds.
6.4 Caveats
All results are based on single runs and are there-
fore not entirely robust. In particular, MERT
tuning of the translation model is known to in-
troduce significant variance in translation perfor-
mance across different runs, and the small differ-
ences in BLEU scores reported in Table 3 are very
likely to lie in that region.
Markov chain convergence also needs further
attention. In absence of complex latent struc-
ture (for the dishes), the chain should mix fairly
quickly, and as attested by Figure 2 it ‘converges’
with respect to the test metric after about 20 sam-
ples, although the log posterior (not shown) had
not converged after 40. The use of a single poste-

rior sample could also be having a negative effect
on results.
7 Future Directions
The first goal will be to get more robust ex-
perimental results, and to scale up to 4-gram
models estimated on all the available monolin-
gual training data. If good performance can be
demonstrated under those conditions, this gen-
eral approach could pass as a viable alternative to
the current Kneser-Ney dominated state-of-the art
setup in MT.
Much of the power of the HPYLM+c model
has not been exploited in this evaluation, in par-
ticular its ability to score unseen compounds con-
sisting of known components. This feature was
71
0 10 20 30 40
Iteration
420
440
460
480
500
520
540
560
Perplexity
KN Λ
ling
HPYLM

HPYLM+c Λ
ling
mKN-baseline
Figure 2: Convergence of test set perplexities.
not active in these evaluations, mostly due to the
current phase of implementation. A second area
of focus is thus to modify the decoder to gen-
erate such unseen compounds in translation hy-
potheses. Given the current low compound recall
rates, this could greatly benefit translation quality.
An informal analysis of the reference translations
in the bilingual test set showed that 991 of the
1406 out-of-vocabulary compounds (out of 2692
OOVs in total) fall into this category of unseen-
but-recognisable compounds.
Ultimately the idea is to apply this modelling
approach to other linguistic phenomena as well.
In particular, the objective is to model instances
of concatenative morphology beyond compound-
ing, with the aim of improving translation into
morphologically rich languages. Complex agree-
ment patterns could be captured by condition-
ing functional morphemes in the target word on
morphemes in the n-gram context, or by stem-
ming context words during back-off. Such ad-
ditional back-off paths can be readily encoded in
the Graphical Pitman-Yor process (Wood and Teh,
2009).
These more complex models may require
longer to train. To this end, I intend to use the

single table per dish approximation (§5) to reduce
training to a single deterministic pass through the
data, conjecturing that this will have little effect
on extrinsic performance.
8 Summary
I have argued for further explorations into the
use of a family of hierarchical Bayesian models
for targeting linguistic phenomena that may not
be captured well by standard n-gram language
models. To ground this investigation, I focused
on German compounds and showed how these
models are an appropriate vehicle for encoding
prior linguistic intuitions about such compounds.
The proposed generative model beats the popu-
lar modified Kneser-Ney model in monolingual
evaluations, and preliminarily achieves small im-
provements in translation from English into Ger-
man. In this translation task, single-token Ger-
man compounds traditionally pose challenges to
translation systems, and preliminary results show
a small increase in the F-score accuracy of com-
pounds in the translation output. Finally, I have
outlined the intended steps for expanding this line
of inquiry into other related linguistic phenomena
and for adapting a translation system to get opti-
mal value out of such improved language models.
Acknowledgements
Thanks goes to my supervisor, Phil Blunsom, for
continued support and advice; to Chris Dyer for
suggesting the focus on German compounds and

supplying a freshly trained compound splitter; to
the Rhodes Trust for financial support; and to the
anonymous reviewers for their helpful feedback.
References
Marco Baroni and Johannes Matiasek. 2002. Pre-
dicting the components of German nominal com-
pounds. In ECAI, pages 470–474.
Andre Berton, Pablo Fetter, and Peter Regel-
Brietzmann. 1996. Compound Words in Large-
Vocabulary German Speech Recognition Systems.
In Proceedings of Fourth International Conference
on Spoken Language Processing. ICSLP ’96, vol-
ume 2, pages 1165–1168. IEEE.
Jeff A Bilmes and Katrin Kirchhoff. 2003. Factored
language models and generalized parallel back-
off. In Proceedings of NAACL-HLT (short papers),
pages 4–6, Stroudsburg, PA, USA. Association for
Computational Linguistics.
Stanley F Chen and Joshua Goodman. 1998. An
Empirical Study of Smoothing Techniques for Lan-
guage Modeling. Technical report.
David Chiang. 2007. Hierarchical Phrase-
Based Translation. Computational Linguistics,
33(2):201–228, June.
Chris Dyer, Adam Lopez, Juri Ganitkevitch, Johnathan
Weese, Ferhan Ture, Phil Blunsom, Hendra Se-
tiawan, Vladimir Eidelman, and Philip Resnik.
2010. cdec: A Decoder, Alignment, and Learning
framework for finite-state and context-free trans-
lation models. In Proceedings of the Association

72
for Computational Linguistics (Demonstration ses-
sion), pages 7–12, Uppsala, Sweden. Association
for Computational Linguistics.
Chris Dyer. 2009. Using a maximum entropy model
to build segmentation lattices for MT. In Proceed-
ings of NAACL, pages 406–414. Association for
Computational Linguistics.
John Goldsmith and Tom Reutter. 1998. Automatic
Collection and Analysis of German Compounds. In
F. Busa F. et al., editor, The Computational Treat-
ment of Nominals, pages 61–69. Universite de Mon-
treal, Canada.
Sharon Goldwater, Thomas L. Griffiths, and Mark
Johnson. 2006. Interpolating Between Types and
Tokens by Estimating Power-Law Generators. In
Advances in Neural Information Processing Sys-
tems, Volume 18.
Songfang Huang and Steve Renals. 2007. Hierarchi-
cal Pitman-Yor Language Models For ASR in Meet-
ings. IEEE ASRU, pages 124–129.
Songfang Huang and Steve Renals. 2009. A paral-
lel training algorithm for hierarchical Pitman-Yor
process language models. In Proceedings of Inter-
speech, volume 9, pages 2695–2698.
Reinhard Kneser and Hermann Ney. 1995. Improved
backing-off for m-gram language modelling. In
Proceedings of the IEEE International Conference
on Acoustics, Speech and SIgnal Processing, pages
181–184.

Philipp Koehn and Kevin Knight. 2003. Empirical
Methods for Compound Splitting. In Proceedings
of EACL, pages 187–193. Association for Compu-
tational Linguistics.
Philipp Koehn, Abhishek Arun, and Hieu Hoang.
2008. Towards better Machine Translation Qual-
ity for the German – English Language Pairs. In
Third Workshop on Statistical Machine Translation,
number June, pages 139–142. Association for Com-
putational Linguistics.
Percy Liang, Ben Taskar, and Dan Klein. 2006. Align-
ment by Agreement. In Proceedings of the Human
Language Technology Conference of the NAACL,
Main Conference, pages 104–111, New York City,
USA, June. Association for Computational Linguis-
tics.
Daichi Mochihashi, Takeshi Yamada, and Naonori
Ueda. 2009. Bayesian unsupervised word seg-
mentation with nested Pitman-Yor language mod-
eling. In Proceedings of the Joint Conference of
the 47th Annual Meeting of the ACL and the 4th In-
ternational Joint Conference on Natural Language
Processing of the AFNLP: Volume 1 - ACL-IJCNLP
’09, pages 100–108, Suntec, Singapore. Associa-
tion for Computational Linguistics.
Radford M Neal. 2003. Slice Sampling. The Annals
of Statistics, 31(3):705–741.
Graham Neubig, Masato Mimura, Shinsuke Mori, and
Tatsuya Kawahara. 2010. Learning a Language
Model from Continuous Speech. In Interspeech,

pages 1053–1056, Chiba, Japan.
Franz Josef Och. 2003. Minimum Error Rate Training
in Statistical Machine Translation. In Proceedings
of ACL, pages 160–167.
Tsuyoshi Okita and Andy Way. 2010. Hierarchical
Pitman-Yor Language Model for Machine Transla-
tion. Proceedings of the International Conference
on Asian Language Processing, pages 245–248.
Kishore Papineni, Salim Roukos, Todd Ward, Wei-
jing Zhu, Thomas J Watson, and Yorktown Heights.
2001. Bleu: A Method for Automatic Evaluation of
Machine Translation. Technical report, IBM.
J Pitman and M. Yor. 1997. The two-parameter
Poisson-Dirichlet distribution derived from a sta-
ble subordinator. The Annals of Probability,
25:855–900.
J. Pitman. 2002. Combinatorial stochastic processes.
Technical report, Department of Statistics, Univer-
sity of California at Berkeley.
Sara Stymne. 2009. A comparison of merging strate-
gies for translation of German compounds. Pro-
ceedings of the 12th Conference of the European
Chapter of the Association for Computational Lin-
guistics: Student Research Workshop, pages 61–69.
Yee Whye Teh. 2006. A hierarchical Bayesian lan-
guage model based on Pitman-Yor processes. In
Proceedings of the 21st International Conference
on Computational Linguistics and the 44th annual
meeting of the ACL, pages 985–992. Association for
Computational Linguistics.

Frank Wood and Yee Whye Teh. 2009. A Hierarchi-
cal Nonparametric Bayesian Approach to Statistical
Language Model Domain Adaptation. In Proceed-
ings of the 12th International Conference on Arti-
ficial Intelligence and Statistics (AISTATS), pages
607–614, Clearwater Beach, Florida, USA.
Mei Yang and Katrin Kirchhoff. 2006. Phrase-based
Backoff Models for Machine Translation of Highly
Inflected Languages. In Proceedings of the EACL,
pages 41–48.
73

×